首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 552 毫秒
1.
C W Mandl  C Kunz    F X Heinz 《Journal of virology》1991,65(8):4070-4077
A poly(A) tail was identified on the 3' end of the prototype tick-borne encephalitis (TBE) virus strain Neudoerfl. This is in contrast to the general lack of poly(A) in the genomic RNAs of mosquito-borne flaviviruses analyzed so far. Analysis of several closely related strains of TBE virus, however, revealed the existence of two different types of 3' noncoding (NC) regions. One type (represented by strain Neudoerfl) is only 114 nucleotides long and carries a 3'-terminal poly(A) structure. This was also found in several TBE virus strains isolated from different geographic regions over a period of almost 30 years. The other type (represented by strain Hypr) is 461 nucleotides long and not polyadenylated. The sequence homology between the two types of TBE virus 3' NC regions terminates at a specific position 81 nucleotides after the stop codon. The second type of 3' NC region more closely resembles the common flavivirus pattern, including the potential for the formation of a 3'-terminal hairpin structure. However, it lacks primary sequence elements that are conserved among other flavivirus genomes.  相似文献   

2.
The flavivirus genome is a positive-strand RNA molecule containing a single long open reading frame flanked by noncoding regions (NCR) that mediate crucial processes of the viral life cycle. The 3′ NCR of tick-borne encephalitis (TBE) virus can be divided into a variable region that is highly heterogeneous in length among strains of TBE virus and in certain cases includes an internal poly(A) tract and a 3′-terminal conserved core element that is believed to fold as a whole into a well-defined secondary structure. We have now investigated the genetic stability of the TBE virus 3′ NCR and its influence on viral growth properties and virulence. We observed spontaneous deletions in the variable region during growth of TBE virus in cell culture and in mice. These deletions varied in size and location but always included the internal poly(A) element of the TBE virus 3′ NCR and never extended into the conserved 3′-terminal core element. Subsequently, we constructed specific deletion mutants by using infectious cDNA clones with the entire variable region and increasing segments of the core element removed. A virus mutant lacking the entire variable region was indistinguishable from wild-type virus with respect to cell culture growth properties and virulence in the mouse model. In contrast, even small extensions of the deletion into the core element led to significant biological effects. Deletions extending to nucleotides 10826, 10847, and 10870 caused distinct attenuation in mice without measurable reduction of cell culture growth properties, which, however, were significantly restricted when the deletion was extended to nucleotide 10919. An even larger deletion (to nucleotide 10994) abolished viral viability. In spite of their high degree of attenuation, these mutants efficiently induced protective immune responses even at low inoculation doses. Thus, 3′-NCR deletions represent a useful technique for achieving stable attenuation of flaviviruses that can be included in the rational design of novel flavivirus live vaccines.The genus Flavivirus (family Flaviviridae) consists of small, enveloped, mainly mosquito- or tick-transmitted viruses with an unsegmented positive-stranded RNA genome (34). Some of these viruses are human pathogens of global medical importance, most notably yellow fever virus, the dengue (DEN) viruses, Japanese encephalitis virus, and tick-borne encephalitis (TBE) virus (22). In spite of the availability of vaccines against several of these viruses, flavivirus infections continue to be a major health problem in many countries of the world. Elucidation of the molecular basis of the pathogenicity of these viruses and identification of specific determinants of virulence are therefore a major focus of flavivirus research.The approximately 11-kb flavivirus genome (for a review, see reference 3) encodes three structural proteins (the capsid protein C, the small membrane protein M, which is formed by proteolytic cleavage from its precursor protein prM, and the large envelope protein E) and seven nonstructural proteins (the glycoprotein NS1, the protease component NS2A, NS2B, the protease/helicase NS3, NS4A, NS4B, and the RNA polymerase NS5). All of the viral proteins are encoded within a single long open reading frame which is flanked by noncoding regions (NCR) believed to carry regulatory elements involved in replication, translation, and packaging of the genome. Molecular analyses of natural low-virulence strains and strains attenuated in vitro by passaging procedures or, more recently, by specific mutagenesis techniques, have shown that genetic determinants that govern the virulence of flaviviruses are located within the coding regions of both structural and nonstructural proteins as well as within the flanking NCRs (2, 21, 26; for reviews, see references 20 and 22). In this study, we focus on the effects of deletions in the 3′ noncoding region (3′ NCR) of TBE virus.TBE virus causes widespread human disease in many European and Asian countries, and its molecular biology has been studied in some detail (29; for a review, see reference 9). The length of the 3′ NCR of TBE virus was previously found to be remarkably heterogeneous even among closely related strains, ranging from approximately 450 to almost 800 nucleotides (31). A more detailed analysis indicated that the 3′ NCR can be divided into a 3′-terminal core element of approximately 340 nucleotides in length and a variable region located between the core element and the open reading frame. The core element is present in all strains investigated so far, and its nucleotide sequence is highly conserved among strains. The entire core element is predicted to fold into a well-defined secondary structure independent of the sequence of the adjacent variable genomic element (27). The variable region is characterized by low sequence conservation, extensive size variability between strains, repetitive sequence elements, and an internal poly(A) tract in certain TBE virus strains (15, 31). Evidence for 3′-NCR size heterogeneity and specific RNA-folding patterns for the 3′-terminal approximately 400 nucleotides have also been observed with several other flaviviruses (5, 24, 25, 33). A similar organization of the 3′ NCRs also appears to be shared by members of the other two genera of the family Flaviviridae, pestiviruses and hepaciviruses (13, 23, 30, 35).Although the functional importance of the flavivirus 3′ NCR is generally acknowledged, the assumed involvement of particular sequence elements in replication, modulation of translation, or packaging is largely hypothetical. Evidence for functionality is so far based mostly on the identification of highly conserved RNA sequence elements or folding patterns by computer techniques. A few studies have provided direct evidence for the binding of protein factors to the stem-loop structure closest to the 3′ terminus (1, 4). Moreover, Men et al. (21) demonstrated that certain deletions introduced into the 3′ NCR of DEN-4 virus resulted in viable mutants with significantly restricted growth properties. By this approach, these researchers were able to identify particular sequences that are required for viability and others that can be deleted without apparent impact on the biology of DEN-4 virus. Studying replicons of Kunjin virus, Khromykh and Westaway (14) found that parts of the 3′ NCR could be deleted or even replaced by a foreign protein expression cassette without loss of replication competence. The 3′-NCR sequences of these flaviviruses, however, exhibit very little homology to the sequences of the tick-borne flaviviruses, which even lack the sequence boxes CS1 and CS2 that are conserved among all mosquito-borne flaviviruses (7, 16).The establishment of an efficient and stable infectious cDNA clone system for TBE virus European subtype prototypic strain Neudoerfl (17) has enabled us to test the functional importance of 3′-NCR sequence elements of this virus by a directed mutagenesis approach. As reported in this communication, spontaneous deletions in the variable region of strain Neudoerfl occur frequently during viral growth in cell culture or in infected animals. This prompted us to construct 3′-NCR deletion mutants of variable lengths to study the influence of these deletions on the biological properties of TBE virus. Our results demonstrate a correlation between the presence of certain RNA sequences or secondary structures and growth properties, viability, and attenuation of the resulting virus mutants. We present several 3′-NCR deletion mutants that are 4 orders of magnitude less virulent than wild-type TBE virus.With regard to vaccine development, the most desirable mutations are ones that are genetically stable and cause significant attenuation but maintain adequate replication properties in cell culture and strong immunogenicity in animals even at low inoculation doses. The evaluation of the TBE virus mutants presented in this article indicates that certain deletions in the 3′ NCR can indeed meet these criteria.  相似文献   

3.
The 3′ noncoding region (NCR) of the negative-strand RNA [3′(−)NCR RNA] of the arterivirus simian hemorrhagic fever virus (SHFV) is 209 nucleotides (nt) in length. Since this 3′ region, designated 3′(−)209, is the site of initiation of full-length positive-strand RNA and is the template for the synthesis of the 5′ leader sequence, which is found on both full-length and subgenomic mRNAs, it is likely to contain cis-acting signals for RNA synthesis and to interact with cellular and viral proteins to form replication complexes. Gel mobility shift assays showed that cellular proteins in MA104 S100 cytoplasmic extracts formed two complexes with the SHFV 3′(−)209 RNA, and results from competition gel mobility shift assays demonstrated that these interactions were specific. Four proteins with molecular masses of 103, 86, 55, and 36 kDa were detected in UV-induced cross-linking assays, and three of these proteins (103, 55, and 36 kDa) were also detected by Northwestern blotting assays. Identical gel mobility shift and UV-induced cross-linking patterns were obtained with uninfected and SHFV-infected extracts, indicating that the four proteins detected are cellular, not viral, proteins. The binding sites for the four cellular proteins were mapped to the region between nt 117 and 184 (68-nt sequence) from the 3′ end of the SHFV negative-strand RNA. This 68-nt sequence was predicted to form two stem-loops, SL4 and SL5. The 3′(−)NCR RNA of another arterivirus, lactate dehydrogenase-elevating virus C (LDV-C), competed with the SHFV 3′(−)209 RNA in competition gel mobility shift assays. UV-induced cross-linking assays showed that four MA104 cellular proteins with the same molecular masses as those that bind to the SHFV 3′(−)209 RNA also bind to the LDV-C 3′(−)NCR RNA and equine arteritis virus 3′(−)NCR RNA. However, each of these viral RNAs also bound to an additional MA104 protein. The binding sites for the MA104 cellular proteins were shown to be located in similar positions in the LDV-C 3′(−)NCR and SHFV 3′(−)209 RNAs. These data suggest that the binding sites for a set of the cellular proteins are conserved in all arterivirus RNAs and that these cell proteins may be utilized as components of viral replication complexes.  相似文献   

4.
The flavivirus genome is a positive-stranded ~11-kb RNA including 5′ and 3′ noncoding regions (NCR) of approximately 100 and 400 to 600 nucleotides (nt), respectively. The 3′ NCR contains adjacent, thermodynamically stable, conserved short and long stem-and-loop structures (the 3′-SL), formed by the 3′-terminal ~100 nt. The nucleotide sequences within the 3′-SL are not well conserved among species. We examined the requirement for the 3′-SL in the context of dengue virus type 2 (DEN2) replication by mutagenesis of an infectious cDNA copy of a DEN2 genome. Genomic full-length RNA was transcribed in vitro and used to transfect monkey kidney cells. A substitution mutation, in which the 3′-terminal 93 nt constituting the wild-type (wt) DEN2 3′-SL sequence were replaced by the 96-nt sequence of the West Nile virus (WN) 3′-SL, was sublethal for virus replication. An analysis of the growth phenotypes of additional mutant viruses derived from RNAs containing DEN2-WN chimeric 3′-SL structures suggested that the wt DEN2 nucleotide sequence forming the bottom half of the long stem and loop in the 3′-SL was required for viability. One 7-bp substitution mutation in this domain resulted in a mutant virus that grew well in monkey kidney cells but was severely restricted in cultured mosquito cells. In contrast, transpositions of and/or substitutions in the wt DEN2 nucleotide sequence in the top half of the long stem and in the short stem and loop were relatively well tolerated, provided the stem-loop secondary structure was conserved.  相似文献   

5.
Enhancement of eukaryotic messenger RNA (mRNA) translation initiation by the 3′ poly(A) tail is mediated through interaction of poly(A)-binding protein with eukaryotic initiation factor (eIF) 4G, bridging the 5′ terminal cap structure. In contrast to cellular mRNA, translation of the uncapped, non-polyadenylated hepatitis C virus (HCV) genome occurs independently of eIF4G and a role for 3′-untranslated sequences in modifying HCV gene expression is controversial. Utilizing cell-based and in vitro translation assays, we show that the HCV 3′-untranslated region (UTR) or a 3′ poly(A) tract of sufficient length interchangeably stimulate translation dependent upon the HCV internal ribosomal entry site (IRES). However, in contrast to cap-dependent translation, the rate of initiation at the HCV IRES was unaffected by 3′-untranslated sequences. Analysis of post-initiation events revealed that the 3′ poly(A) tract and HCV 3′-UTR improve translation efficiency by enabling termination and possibly ribosome recycling for successive rounds of translation.  相似文献   

6.
The 3′ end region of foot-and-mouth disease virus (FMDV) consists of two distinct elements, a 90 nt untranslated region (3′-NCR) and a poly(A) tract. Removal of either the poly(A) tract or both the 3′-NCR and the poly(A) tract abrogated infectivity in susceptible cells in the context of a full-length cDNA clone. We have addressed the question of whether the impairment of RNA infectivity is related to defects at the translation level using a double approach. First, compared to the full-length viral RNA, removal of the 3′ sequences reduced the efficiency of translation in vitro. Secondly, a stimulatory effect of the 3′ end sequences on IRES-dependent translation was found in vivo using bicistronic constructs. RNAs carrying the FMDV 3′ end sequences linked to the second cistron showed a significant stimulation of IRES-dependent translation, whereas cap-dependent translation was not affected. Remarkably, IRES-dependent stimulation exerted by the poly(A) tract or the 3′-NCR seems to be the result of two separate events, as the 3′-NCR alone enhanced IRES activity on its own. Under conditions of FMDV Lb protease-induced translation shut-off, the stimulation of IRES activity reached values above 6-fold in living cells. A northern blot analysis indicated that IRES stimulation was not the consequence of a change in the stability of the bicistronic RNA produced in transfected cells. Analysis of the RNA-binding proteins interacting with a mixture of 3′ end and IRES probes showed an additive pattern. Altogether, our results strongly suggest that individual signals in the viral 3′ end ensure stimulation of FMDV translation.  相似文献   

7.
Genomes and antigenomes of many positive-strand RNA viruses contain 3′-poly(A) and 5′-poly(U) tracts, respectively, serving as mutual templates. Mechanism(s) controlling the length of these homopolymeric stretches are not well understood. Here, we show that in coxsackievirus B3 (CVB3) and three other enteroviruses the poly(A) tract is ~80–90 and the poly(U) tract is ~20 nt-long. Mutagenesis analysis indicate that the length of the CVB3 3′-poly(A) is determined by the oriR, a cis-element in the 3′-noncoding region of viral RNA. In contrast, while mutations of the oriR inhibit initiation of (−) RNA synthesis, they do not affect the 5′-poly(U) length. Poly(A)-lacking genomes are able to acquire genetically unstable AU-rich poly(A)-terminated 3′-tails, which may be generated by a mechanism distinct from the cognate viral RNA polyadenylation. The aberrant tails ensure only inefficient replication. The possibility of RNA replication independent of oriR and poly(A) demonstrate that highly debilitated viruses are able to survive by utilizing ‘emergence’, perhaps atavistic, mechanisms.  相似文献   

8.

Background

With the development of new specific inhibitors of hepatitis C virus (HCV) enzymes and functions that may yield different antiviral responses and resistance profiles according to the HCV subtype, correct HCV genotype 1 subtype identification is mandatory in clinical trials for stratification and interpretation purposes and will likely become necessary in future clinical practice. The goal of this study was to identify the appropriate molecular tool(s) for accurate HCV genotype 1 subtype determination.

Methodology/Principal Findings

A large cohort of 500 treatment-naïve patients eligible for HCV drug trials and infected with either subtype 1a or 1b was studied. Methods based on the sole analysis of the 5′ non-coding region (5′NCR) by sequence analysis or reverse hybridization failed to correctly identify HCV subtype 1a in 22.8%–29.5% of cases, and HCV subtype 1b in 9.5%–8.7% of cases. Natural polymorphisms at positions 107, 204 and/or 243 were responsible for mis-subtyping with these methods. A real-time PCR method using genotype- and subtype-specific primers and probes located in both the 5′NCR and the NS5B-coding region failed to correctly identify HCV genotype 1 subtype in approximately 10% of cases. The second-generation line probe assay, a reverse hybridization assay that uses probes targeting both the 5′NCR and core-coding region, correctly identified HCV subtypes 1a and 1b in more than 99% of cases.

Conclusions/Significance

In the context of new HCV drug development, HCV genotyping methods based on the exclusive analysis of the 5′NCR should be avoided. The second-generation line probe assay is currently the best commercial assay for determination of HCV genotype 1 subtypes 1a and 1b in clinical trials and practice.  相似文献   

9.
Utilization of internal ribosome entry segment (IRES) structures in the 5′ noncoding region (5′NCR) of picornavirus RNAs for initiation of translation requires a number of host cell factors whose distribution may vary in different cells and whose requirement may vary for different picornaviruses. We have examined the requirement of the cellular protein poly(rC) binding protein 2 (PCBP2) for hepatitis A virus (HAV) RNA translation. PCBP2 has recently been identified as a factor required for translation and replication of poliovirus (PV) RNA. PCBP2 was shown to be present in FRhK-4 cells, which are permissive for growth of HAV, as it is in HeLa cells, which support translation of HAV RNA but which have not been reported to host replication of the virus. Competition RNA mobility shift assays showed that the 5′NCR of HAV RNA competed for binding of PCBP2 with a probe representing stem-loop IV of the PV 5′NCR. The binding site on HAV RNA was mapped to nucleotides 1 to 157, which includes a pyrimidine-rich sequence. HeLa cell extracts that had been depleted of PCBP2 by passage over a PV stem-loop IV RNA affinity column supported only low levels of HAV RNA translation. Translation activity was restored upon addition of recombinant PCBP2 to the depleted extract. Removal of the 5′-terminal 138 nucleotides of the HAV RNA, or removal of the entire IRES, eliminated the dependence of HAV RNA translation on PCBP2.  相似文献   

10.
Alphaviruses are mosquito-transmitted RNA viruses that cause important diseases in both humans and livestock. Sindbis virus (SIN), the type species of the alphavirus genus, carries a 11.7-kb positive-sense RNA genome which is capped at its 5′ end and polyadenylated at its 3′ end. The 3′ nontranslated region (3′NTR) of the SIN genome carries many AU-rich motifs, including a 19-nucleotide (nt) conserved element (3′CSE) and a poly(A) tail. This 3′CSE and the adjoining poly(A) tail are believed to regulate the synthesis of negative-sense RNA and genome replication in vivo. We have recently demonstrated that the SIN genome lacking the poly(A) tail was infectious and that de novo polyadenylation could occur in vivo (K. R. Hill, M. Hajjou, J. Hu, and R. Raju, J. Virol. 71:2693–2704, 1997). Here, we demonstrate that the 3′-terminal 29-nt region of the SIN genome carries a signal for possible cytoplasmic polyadenylation. To further investigate the polyadenylation signals within the 3′NTR, we generated a battery of mutant genomes with mutations in the 3′NTR and tested their ability to generate infectious virus and undergo 3′ polyadenylation in vivo. Engineered SIN genomes with terminal deletions within the 19-nt 3′CSE were infectious and regained their poly(A) tail. Also, a SIN genome carrying the poly(A) tail but lacking a part or the entire 19-nt 3′CSE was also infectious. Sequence analysis of viruses generated from these engineered SIN genomes demonstrated the addition of a variety of AU-rich sequence motifs just adjacent to the poly(A) tail. The addition of AU-rich motifs to the mutant SIN genomes appears to require the presence of a significant portion of the 3′NTR. These results indicate the ability of alphavirus RNAs to undergo 3′ repair and the existence of a pathway for the addition of AU-rich sequences and a poly(A) tail to their 3′ end in the infected host cell. Most importantly, these results indicate the ability of alphavirus replication machinery to use a multitude of AU-rich RNA sequences abutted by a poly(A) motif as promoters for negative-sense RNA synthesis and genome replication in vivo. The possible roles of cytoplasmic polyadenylation machinery, terminal transferase-like enzymes, and the viral polymerase in the terminal repair processes are discussed.  相似文献   

11.
DNA-based antisense oligonucleotides (ASOs) elicit cleavage of the targeted RNA by the endoribonuclease RNase H1, whereas siRNAs mediate cleavage through the RNAi pathway. To determine the fates of the cleaved RNA in cells, we lowered the levels of the factors involved in RNA surveillance prior to treating cells with ASOs or siRNA and analyzed cleavage products by RACE. The cytoplasmic 5′ to 3′ exoribonuclease XRN1 was responsible for the degradation of the downstream cleavage products generated by ASOs or siRNA targeting mRNAs. In contrast, downstream cleavage products generated by ASOs targeting nuclear long non-coding RNA Malat 1 and pre-mRNA were degraded by nuclear XRN2. The downstream cleavage products did not appear to be degraded in the 3′ to 5′ direction as the majority of these products contained intact poly(A) tails and were bound by the poly(A) binding protein. The upstream cleavage products of Malat1 were degraded in the 3′ to 5′ direction by the exosome complex containing the nuclear exoribonuclease Dis3. The exosome complex containing Dis3 or cytoplasmic Dis3L1 degraded mRNA upstream cleavage products, which were not bound by the 5′-cap binding complex and, consequently, were susceptible to degradation in the 5′ to 3′ direction by the XRN exoribonucleases.  相似文献   

12.
You S  Rice CM 《Journal of virology》2008,82(1):184-195
The hepatitis C virus (HCV) genomic RNA possesses conserved structural elements that are essential for its replication. The 3′ nontranslated region (NTR) contains several of these elements: a variable region, the poly(U/UC) tract, and a highly conserved 3′ X tail, consisting of stem-loop 1 (SL1), SL2, and SL3. Studies of drug-selected, cell culture-adapted subgenomic replicons have indicated that an RNA element within the NS5B coding region, 5BSL3.2, forms a functional kissing-loop tertiary structure with part of the 3′ NTR, 3′ SL2. Recent advances now allow the efficient propagation of unadapted HCV genomes in the context of a complete infectious life cycle (HCV cell culture [HCVcc]). Using this system, we determine that the kissing-loop interaction between 5BSL3.2 and 3′ SL2 is required for replication in the genotype 2a HCVcc context. Remarkably, the overall integrity of the 5BSL3 cruciform is not an absolute requirement for the kissing-loop interaction, suggesting a model in which trans-acting factor(s) that stabilize this interaction may interact initially with the 3′ X tail rather than 5BSL3. The length and composition of the poly(U/UC) tract were also critical determinants of HCVcc replication, with a length of 33 consecutive U residues required for maximal RNA amplification. Interrupting the U homopolymer with C residues was deleterious, implicating a trans-acting factor with a preference for U over mixed pyrimidine nucleotides. Finally, we show that both the poly(U) and kissing-loop RNA elements can function outside of their normal genome contexts. This suggests that the poly(U/UC) tract does not function simply as an unstructured spacer to position the kissing-loop elements.  相似文献   

13.
Japanese encephalitis virus (JEV) has a single-stranded, positive-sense RNA genome containing a single open reading frame flanked by the 5′- and 3′-non-coding regions (NCRs). The virus genome replicates via a negative-sense RNA intermediate. The NCRs and their complementary sequences in the negative-sense RNA are the sites for assembly of the RNA replicase complex thereby regulating the RNA synthesis and virus replication. In this study, we show that the 55-kDa polypyrimidine tract-binding protein (PTB) interacts in vitro with both the 5′-NCR of the positive-sense genomic RNA - 5NCR(+), and its complementary sequence in the negative-sense replication intermediate RNA - 3NCR(-). The interaction of viral RNA with PTB was validated in infected cells by JEV RNA co-immunoprecipitation and JEV RNA-PTB colocalization experiments. Interestingly, we observed phosphorylation-coupled translocation of nuclear PTB to cytoplasmic foci that co-localized with JEV RNA early during JEV infection. Our studies employing the PTB silencing and over-expression in cultured cells established an inhibitory role of PTB in JEV replication. Using RNA-protein binding assay we show that PTB competitively inhibits association of JEV 3NCR(-) RNA with viral RNA-dependent RNA polymerase (NS5 protein), an event required for the synthesis of the plus-sense genomic RNA. cAMP is known to promote the Protein kinase A (PKA)-mediated PTB phosphorylation. We show that cells treated with a cAMP analogue had an enhanced level of phosphorylated PTB in the cytoplasm and a significantly suppressed JEV replication. Data presented here show a novel, cAMP-induced, PTB-mediated, innate host response that could effectively suppress JEV replication in mammalian cells.  相似文献   

14.
Alternative polyadenylation leads to mRNAs with variable 3′ ends. Since a 3′-untranslated region (3′-UTR) often contains cis elements that impact stability or localization of mRNA or translation, selection of poly(A) sites in a 3′-UTR is regulated in mammalian cells. However, the molecular basis for alternative poly(A) site selection within a 3′-UTR has been unclear. Here we show involvement of cleavage factor Im (CFIm) in poly(A) site selection within a 3′-UTR. CFIm is a heterodimeric 3′ end-processing complex, which functions to assemble other processing factors on pre-mRNA in vitro. We knocked down 25 kDa subunit of CFIm (CFIm25) in HeLa cells and analyzed alternative poly(A) site selection of TIMP-2, syndecan2, ERCC6 and DHFR genes by northern blotting. We observed changes in the distribution of mRNAs in CFIm25 depleted cells, suggesting a role for CFIm in alternative poly(A) site selection. Furthermore, tissue specific analysis demonstrated that the CFIm25 gene gave rise to 1.1, 2.0 and 4.6 kb mRNAs. The 4.6 kb mRNA was ubiquitously expressed, while the 1.1 and 2.0 kb mRNAs were expressed in a tissue specific manner. We found three likely poly(A) sites in the CFIm25 3′-UTR, suggesting alternative polyadenylation. Our results indicate that alternative poly(A) site selection is a well-regulated process in vivo.  相似文献   

15.
16.
Sewage surveillance in seven Italian cities between 2005 and 2008, after the introduction of inactivated poliovirus vaccination (IPV) in 2002, showed rare polioviruses, none that were wild-type or circulating vaccine-derived poliovirus (cVDPV), and many other enteroviruses among 1,392 samples analyzed. Two of five polioviruses (PV) detected were Sabin-like PV2 and three PV3, based on enzyme-linked immunosorbent assay (ELISA) and PCR results. Neurovirulence-related mutations were found in the 5′ noncoding region (5′NCR) of all strains and, for a PV2, also in VP1 region 143 (Ile > Thr). Intertypic recombination in the 3D region was detected in a second PV2 (Sabin 2/Sabin 1) and a PV3 (Sabin 3/Sabin 2). The low mutation rate in VP1 for all PVs suggests limited interhuman virus passages, consistent with efficient polio immunization in Italy. Nonetheless, these findings highlight the risk of wild or Sabin poliovirus reintroduction from abroad. Non-polio enteroviruses (NPEVs) were detected, 448 of which were coxsackievirus B (CVB) and 294 of which were echoviruses (Echo). Fifty-six NPEVs failing serological typing were characterized by sequencing the VP1 region (nucleotides [nt] 2628 to 2976). A total of 448 CVB and 294 Echo strains were identified; among those strains, CVB2, CVB5, and Echo 11 predominated. Environmental CVB5 and CVB2 strains from this study showed high sequence identity with GenBank global strains. The high similarity between environmental NPEVs and clinical strains from the same areas of Italy and the same periods indicates that environmental strains reflect the viruses circulating in the population and highlights the potential risk of inefficient wastewater treatments. This study confirmed that sewage surveillance can be more sensitive than acute flaccid paralysis (AFP) surveillance in monitoring silent poliovirus circulation in the population as well as the suitability of molecular approaches to enterovirus typing.  相似文献   

17.
In this paper, we describe a production procedure of the one-to-one double helical complex of poly(dG)–poly(dC), characterized by a well-defined length (up to 10 kb) and narrow size distribution of molecules. Direct evidence of strands slippage during poly(dG)–poly(dC) synthesis by Klenow exo fragment of polymerase I is obtained by fluorescence resonance energy transfer (FRET). We show that the polymer extension results in an increase in the separation distance between fluorescent dyes attached to 5′ ends of the strands in time and, as a result, losing communication between the dyes via FRET. Analysis of the products of the early steps of the synthesis by high-performance liquid chromatography and mass spectroscopy suggest that only one nucleotide is added to each of the strand composing poly(dG)–poly(dC) in the elementary step of the polymer extension. We show that proper pairing of a base at the 3′ end of the primer strand with a base in sequence of the template strand is required for initiation of the synthesis. If the 3′ end nucleotide in either poly(dG) or poly(dC) strand is substituted for A, the polymer does not grow. Introduction of the T-nucleotide into the complementary strand to permit pairing with A-nucleotide results in the restoration of the synthesis. The data reported here correspond with a slippage model of replication, which includes the formation of loops on the 3′ ends of both strands composing poly(dG)–poly(dC) and their migration over long-molecular distances (μm) to 5′ ends of the strands.  相似文献   

18.
Porcine sapelovirus (PSV), a species of the genus Sapelovirus within the family Picornaviridae, is associated with diarrhea, pneumonia, severe neurological disorders, and reproductive failure in pigs. However, the structural features of the complete PSV genome remain largely unknown. To analyze the structural features of PSV genomes, the full-length nucleotide sequences of three Korean PSV strains were determined and analyzed using bioinformatic techniques in comparison with other known PSV strains. The Korean PSV genomes ranged from 7,542 to 7,566 nucleotides excluding the 3′ poly(A) tail, and showed the typical picornavirus genome organization; 5′untranslated region (UTR)-L-VP4-VP2-VP3-VP1-2A-2B-2C-3A-3B-3C-3D-3′UTR. Three distinct cis-active RNA elements, the internal ribosome entry site (IRES) in the 5′UTR, a cis-replication element (CRE) in the 2C coding region and 3′UTR were identified and their structures were predicted. Interestingly, the structural features of the CRE and 3′UTR were different between PSV strains. The availability of these first complete genome sequences for PSV strains will facilitate future investigations of the molecular pathogenesis and evolutionary characteristics of PSV.  相似文献   

19.
The 3′ untranslated genome region (UTR) of arthropod-borne viruses is characterized by enriched direct repeats (DRs) and stem-loop structures. Despite many years of theoretical and experimental study, on-going positive selection on the 3′UTR had never been observed in ‘real-time,’ and the role of the arbovirus 3′UTR remains poorly understood. We observed a lineage-specific 3′UTR sequence pattern in all available Asian lineage of the mosquito-borne alphavirus, chikungunya virus (CHIKV) (1958–2009), including complicated mutation and duplication patterns of the long DRs. Given that a longer genome is usually associated with less efficient replication, we hypothesized that the fixation of these genetic changes in the Asian lineage 3′UTR was due to their beneficial effects on adaptation to vectors or hosts. Using reverse genetic methods, we examined the functional importance of each direct repeat. Our results suggest that adaptation to mosquitoes, rather than to mammalian hosts, is a major evolutionary force on the CHIKV 3′UTR. Surprisingly, the Asian 3′UTR appeared to be inferior to its predicted ancestral sequence for replication in both mammals and mosquitoes, suggesting that its fixation in Asia was not a result of directional selection. Rather, it may have resulted from a population bottleneck during its introduction from Africa to Asia. We propose that this introduction of a 3′UTR with deletions led to genetic drift and compensatory mutations associated with the loss of structural/functional constraints, followed by two independent beneficial duplications and fixation due to positive selection. Our results provide further evidence that the limited epidemic potential of the Asian CHIKV strains resulted from founder effects that reduced its fitness for efficient transmission by mosquitoes there.  相似文献   

20.
The complete nucleotide sequence of a novel enteric virus, Aichi virus, associated with nonbacterial acute gastroenteritis in humans was determined. The Aichi virus genome proved to be a single-stranded positive-sense RNA molecule with 8,251 bases excluding a poly(A) tail; it contains a large open reading frame with 7,302 nucleotides that encodes a potential polyprotein precursor of 2,433 amino acids. The genome contains a 5′ nontranslated region (NTR) with 712 bases and a 3′ NTR with 240 bases followed by a poly(A) tail. The structure of the genome, VPg–5′ NTR–leader protein–structural proteins–nonstructural proteins–3′ NTR–poly(A), was found to be typical of a picornavirus. The VP0-VP3 and VP3-VP1 cleavage sites were determined to be Q-H and Q-T, respectively, by N-terminal amino acid sequence analyses using purified virion proteins. Possible cleavage sites, Q-G, Q-A, and Q-S, which cleave P2 and P3 polyproteins were found to be similar to those of picornaviruses. A dendrogram based on 3Dpol proteins indicated that Aichi virus is genetically distinct from the known six genera of picornaviruses including entero-, rhino-, cardio-, aphtho-, and hepatovirus and echovirus 22. Considering this together with other properties of the virus (T. Yamashita, S. Kobayashi, K. Sakae, S. Nakata, S. Chiba, Y. Ishihara, and S. Isomura, J. Infect. Dis. 164:954–957, 1991), we propose that Aichi virus be regarded as a new genus of the family Picornaviridae.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号